Thermo-economic analysis and optimization of a combined cooling and power (CCP) system for engine waste heat recovery

Thermo-economic analysis and optimization of a combined cooling and power (CCP) system for engine waste heat recovery

Energy Conversion and Management 128 (2016) 303–316 Contents lists available at ScienceDirect Energy Conversion and Management journal homepage: www...

1MB Sizes 0 Downloads 55 Views

Energy Conversion and Management 128 (2016) 303–316

Contents lists available at ScienceDirect

Energy Conversion and Management journal homepage: www.elsevier.com/locate/enconman

Thermo-economic analysis and optimization of a combined cooling and power (CCP) system for engine waste heat recovery Jiaxi Xia, Jiangfeng Wang ⇑, Juwei Lou, Pan Zhao, Yiping Dai Institute of Turbomachinery, State Key Laboratory of Multiphase Flow in Power Engineering, School of Energy and Power Engineering, Xi’an Jiaotong University, Xi’an 710049, China

a r t i c l e

i n f o

Article history: Received 29 June 2016 Received in revised form 28 August 2016 Accepted 27 September 2016

Keywords: Internal combustion engine Waste heat recovery Brayton cycle Organic Rankine cycle Ejector refrigeration cycle Optimization

a b s t r a c t A combined cooling and power (CCP) system is developed, which comprises a CO2 Brayton cycle (BC), an organic Rankine cycle (ORC) and an ejector refrigeration cycle for the cascade utilization of waste heat from an internal combustion engine. By establishing mathematical model to simulate the overall system, thermodynamic analysis and exergoeconomic analysis are conducted to examine the effects of five key parameters including the compressor pressure ratio, the compressor inlet temperature, the BC turbine inlet temperature, the ORC turbine inlet pressure and the ejector primary flow pressure on system performance. What’s more, a single-objective optimization by means of genetic algorithm (GA) is carried out to search the optimal system performance from viewpoint of exergoeconomic. Results show that the increases of the BC turbine inlet temperature, the ORC turbine inlet pressure and the ejector primary flow pressure are benefit to both thermodynamic and exergoeconimic performances of the CCP system. However, the rises in compressor pressure ratio and compressor inlet temperature will lead to worse system performances. By the single-objective optimization, the lowest average cost per unit of exergy product for the overall system is obtained. Ó 2016 Elsevier Ltd. All rights reserved.

1. Introduction With the development of world economy and industry, the energy shortage and environmental problems caused by fossil fuel consumption have attracted more and more attention. Internal combustion engines (ICES) which act as the major source of motive power, consume a large proportion of petroleum resources in the world. It is reported that only about one-third of the fuel energy in the internal combustion engines is converted into mechanical work, the remaining energy is mainly wasted by rejecting heat to the environment through the exhaust and the coolant [1]. Therefore, it would be of great significance to recover the waste heat effectively from internal combustion engines. The technique of organic Rankine cycle (ORC) has been proven to be a potential method to convert the engine waste heat into power, since it presents some advantages of operation flexibility and desirable efficiency [2,3]. Much work has been carried out on the ORC configurations for the engine waste heat recovery. Vaja and Gambarotta [4] examined three different ORC setups, namely a simple cycle only recovering engine exhaust gases, a simple cycle recovering both exhaust gases and engine cooling water with a ⇑ Corresponding author. E-mail address: [email protected] (J. Wang). http://dx.doi.org/10.1016/j.enconman.2016.09.086 0196-8904/Ó 2016 Elsevier Ltd. All rights reserved.

preheater and a cycle with regeneration. Tahani et al. [5] introduced two different kinds of ORC including the preheat and twostage configurations for the waste heat recovery of engine exhaust gases and coolant. He et al. [6] developed a combined thermodynamic system which contains two cycles: an ORC for the recovery of waste heat from high temperature exhaust gases and the lubricant, a Kalina cycle for the recovery of waste heat from cooling water. Wang et al. [7] studied the thermal performance of a dual-loop ORC coupled with a gasoline engine. Kim et al. [8] proposed a highly efficient single-loop ORC for the waste heat recovery of exhaust gases and coolant from a gasoline vehicle. Since the choice of working fluids had a great influence on thermodynamic performance of an ORC [9], other researchers had put their focuses on working fluid selection [10–13], considering both pure working fluids and zeotropic mixtures for ORCs used in waste heat recovery of ICEs. Regarding all these studies mentioned above, the heat exchanges between exhaust gases and organic working fluids were conducted directly, that may cause decompositions of organic working fluids, due to the high temperature of exhaust gases (about 450–600 °C) and the low decomposition temperatures of organic working fluids (about 200–300 °C). In order to avoid this issue, an intermediate loop with thermal oil is placed between the exhaust gases and the ORC [14,15]. Although the stability

304

J. Xia et al. / Energy Conversion and Management 128 (2016) 303–316

Nomenclature A Bo C CRF CEPCI c cp D E f G h ieff L M n Nu P Pr Pt pr Q Qvs qm r s T U V v W X x Z

area, m2 boiling number cost rate, $ year1 capital recovery factor chemical engineering plant cost index average cost per unit of exergy, $ (MW h)1 specific heat, kJ kg K1 diameter, m exergy flow rate, kJ s1 friction factor mass velocity, kg m2 s1 enthalpy, kJ kg1 interest rate length, m mass flow rate, kg s1 lifetime, year Nusselt number pressure, MPa Prandtl number center distance between tubes, m reduced pressure heat transfer rate, kW volumetric steam flow, m3 s1 average imposed wall heat flux, W m2 enthalpy of vaporization, kJ kg1 entropy, kJ kg1 K1 temperature, K overall heat transfer coefficient, W m2 K1 volume, m3 velocity, m s1 power, kW size or capacity parameter vapor quality annually levelized cost value, $ year1

Greek symbol a convection heat transfer coefficient, W m2 K1 g efficiency, % k thickness, m q density, kg m3 l dynamic viscosity, m2 s1 p compressor pressure ratio

0 BC BM bt cf cnd comp cool cond1 cond2 D eq es evp exg evap F gh he hf in ip L l M m ORC ot out P pc pump pump1 pump2 sp sep turb v w wbt wot

ambient state Brayton cycle bare module BC turbine cold fluid condensation compressor refrigeration capacity output condenser 1 condenser 2 destruction equipment equivalent diameter evaporation exergy evaporator fuel gas heater heat exchanger hot fluid inside inlet pipe loss liquid material mean organic Rankine cycle ORC turbine outside product; pressure precooler pump pump 1 pump 2 single-phase separator turbine vapor tube wall power produced by BC turbine power produced by ORC turbine

Subscript 1–29 state points a–e state points

and safety of the system could be improved through this method, the great amount of high-temperature exhaust gases heat is not exploited at all. Several studies has been performed on exploring novel dual-loop systems combined ORCs with other thermodynamic cycles for engine waste heat recovery. Mliller et al. [16] developed a system that combined ORC with thermoelectric conversion. Through a thermoelectric generator (TEG), the high temperature exhaust heat was converted into power, and the working fluid of ORC was preheated. But the application of this combined system is constrained due to the low energy conversion capacity of TEG [17]. Zhang et al. [18] analyzed the characteristics of a dual loop ORC system which was combined with a vehicular light-duty diesel engine. Yang et al. [19] explored system performance of the dual loop ORC system used for diesel engine waste heat recovery under various operating conditions. Choi and Kim

[20] presented a dual-loop power generation system including an upper trilateral cycle with water and a bottoming ORC. The former utilized the waste heat from the high-temperature exhaust gases discharged by a marine engine, and the latter reused the turbine exhaust heat of the upper cycle and the low-temperature exhaust gases waste heat. Shu et al. [21] designed a novel dual-loop ORC system which consists of a high-temperature (HT) loop and a low-temperature (LT). The HT loop was a steam Rankine cycle used for the waste heat recovery of the high-temperature part of exhaust gases, while the LT loop was an organic Ranking cycle recovering the coolant heat, the turbine exhaust heat from HT loop and the waste heat from low-temperature part of exhaust gases. With regard to the same dual-loop ORC system, Song and Gu [22] considered the exploiting of wet steam expansion by applying a screw expander to the HT loop, and investigated the effect of the

305

J. Xia et al. / Energy Conversion and Management 128 (2016) 303–316

HT loop parameters on LT loop. Zhou et al. [23] studied the dualloop ORC with zeotropic mixtures as the working fluid in the low temperature loop. For the sake of satisfying the diverse consumers’ demands, the combined cooling and power (CCP) concept is also introduced for the engine waste heat recovery by a number of researchers. Liang et al. [24] developed a power and cooling cogeneration system which included an ORC and an absorption refrigeration cycle to utilize the exhaust gases heat of a marine engine. Wang [25] integrated the transcritical CO2 refrigeration cycle with the supercritical CO2 cycle to recover the waste heat of an ICE, producing both cooling and power energy. However, for the dual-loop ORCs, most of the recent studies only considered the steam Rankine cycle as the HT loop for the heat recovery of high-temperature exhaust gases, neglecting the Brayton cycle, which has been proven to be suitable for recovering waste heat from high-temperature heat source, due to its highly efficiency, small structure and simply layout [26–28]. Even though Zhang et al. [17] have done some work about this, the waste heat recovery of the engine coolant was not taken into account, and the detailed parameter analysis including both HT-loop and LT-loop were not given. Meanwhile, in order to avoid the issue of the mismatching mass flow rate which is caused by the setting in series of the preheater utilized engine coolant heat and the evaporator utilized the residual heat of HT loop, only a part of engine coolant heat could be recovered [4,5,21]. What’s more, few studies conducted the performance from the viewpoint of exergoeconomic for the system used for engine waste heat recovery. Referring to the CCP systems, little information has been published concerning their applications to recover both exhaust gases heat and coolant heat from engines. In this study, an organic Rankine cycle coupled with a CO2 Brayton cycle is considered. In order to utilize the energy source more efficiently, an ejector refrigeration cycle is added as well, since it can utilize low-grade heat sources and has advantages of lower costs, higher reliability and simpler operation [29]. Therefore, a combined cooling and power (CCP) system is designed to recover the waste heat of an engine. The CO2 Brayton cycle is used to recover the waste heat of the high temperature exhaust gases, the organic Rankine cycle with a separator is utilized to make full use of the coolant heat and the residual heat of Brayton cycle, while the ejector refrigeration cycle is applied to reuse the waste heat of the low temperature exhaust gases. The effects of five key parameters on thermodynamic and exergoeconimic performances of the CCP system are examined. Furthermore, a single-objective optimization by means of genetic algorithm is carried out to obtain a better system performance. 2. System description The CCP system designed for this study is shown in Fig. 1. It consists of a CO2 Brayton cycle (BC), an organic Rankine cycle (ORC), and an ejector refrigeration cycle, which produces both power and refrigeration simultaneously. The high temperature exhaust gases are firstly delivered to the gas heater to drive the BC. Cooled CO2 from the precooler is compressed to supercritical state by compressor, and then passes through the gas heater to absorb heat, becoming high temperature and high pressure vapor. The supercritical CO2 enters to BC turbine where it expands to produce power, and the CO2 exhaust is cooled by the vapor generator 1 and precooler in turn to complete the CO2 Brayton cycle. For the sake of recovering residual heat of CO2 exhaust from the BC turbine, the bottoming loop (ORC) is coupled to the upper loop (BC) via vapor generator 1. Liquid organic working fluid from condenser 1 is pressured by pump 1 and delivered to the preheater, where it is heated by jacket water and becomes two-phase state. The two-phase working fluid is then delivered to separator and

separated into saturated vapor and saturated liquid. After the separation process, saturated liquid goes through vapor generator 1 to absorb heat from CO2 exhaust, producing superheated vapor. Then the superheated vapor is mixed with the saturated vapor from separator, and expands through the ORC turbine for power generation. At last, the exhaust from the ORC turbine is condensed to be liquid. The ejector refrigeration cycle is used to recover the residual heat energy of exhaust gases after releasing heat through the gas heater. Liquid working fluid is divided into two parts after the condensation process in condenser 2. One part of working fluid enters pump 2 to be pumped to vapor generator 2, producing superheated vapor by absorbing heat from exhaust gas. The other part of working fluid passes through the throttle valve, being vaporized to vapor state in evaporator by absorbing heat from the cooling side, producing cooling capacity. After that, these two parts of working fluids are mixed in the ejector, and then enter condenser 2 to be condensed to liquid. Isobutane is selected as the working fluid for Organic Rankine cycle and ejector refrigeration cycle, since it is suitable for hightemperature ORCs [30,31] and has good performance in thermodynamic and environmental fields [32]. In addition, water at ambient temperature is chosen as the coolant in the precooler and condensers.

3. Mathematical models and performance criteria In order to simplify the simulation of the system, several assumptions are employed as follows: (1) The system reaches a steady state. (2) The pressure losses in pipes are neglected and there is a 2% pressure drop in each heat exchanger. (3) The heat losses in each component are not taken into account. (4) The working fluid at the condenser outlet is saturated liquid. And the state of evaporator outlet is saturated vapor. (5) The isentropic efficiencies of turbines and compressor are 80%, and those of pumps are set to be 70%. (6) The exhaust temperature at the outlet of vapor generator 2 is confined to be over 110 °C, in order to avoid the lowtemperature corrosion [33]. 3.1. Thermodynamic analysis 3.1.1. Energy analysis The mathematical model of energy analysis for each component is established based on the principles of mass and energy conservation, the equations are given by

X

X

M in ¼ Q

X

X

ð1Þ

M out



X

M out  hout 

X

M in  hin

ð2Þ

The detailed energy balance equations for the CCP system components are listed in Table 1. The net power output of Brayton cycle is expressed as

W BC ¼ W bt  W comp

ð3Þ

The net power output of organic Rankine is calculated as

W ORC ¼ W ot  W pump1

ð4Þ

The net power output of system is given by

W net ¼ W bt þ W ot  W comp  W pump1  W pump2

ð5Þ

The cooling capacity output is

Q cool ¼ M 28  ðh28  h29 Þ

ð6Þ

306

J. Xia et al. / Energy Conversion and Management 128 (2016) 303–316

Fig. 1. Schematic diagram of the CCP system.

3.2. Exergoeconomic analysis

Table 1 Energy relations for each component. Component

Energy relations

Gas heater BC turbine Vapor generator 1 Precooler Compressor ORC turbine Condenser 1 Pump1 Preheater Separator Vapor generator 2 Ejector Evaporator Throttle valve

Ma  (ha  hb) = M1  (h1  h5) Wbt = M1  (h1  h2) = M1  (h1  h2s)  gbt M2  (h2  h3) = M6  (h6  h13) M3  (h3  h4) = M22  (h23  h22) Wcomp = M4  (h5  h4) = M4  (h5s  h4)/gcomp Wot = M7  (h7  h8) = M7  (h7  h8s)  got M8  (h8  h9) = M24  (h25  h24) Wpump1 = M9  (h10  h9) = M9  (h10s  h9)/gpump1 Md  (hd  he) = M10  (h11  h10) M11  h11 = M12  h12 + M13  h13 Mb  (hb  hc) = M14  (h14  h21) M15  h15 = M14  h14 + M19  h19 M19  (h19  h18) = M28  (h28  h29) h18 = h17

3.2.1. Size of main equipments 3.2.1.1. Areas of heat exchangers. In this study, all heat exchangers are set as shell-and-tube type. Single-phase heat transfer process occurs in the gas heater and the precooler, two-phase heat transfer process takes place in the evaporator, while both single-phase and two-phase heat transfer processes are conducted in the vapor generators and condensers. As the thermodynamic properties of the working fluid will undergo changes in heat convection heat transfer process, this process is discretized to many subsections for each heat exchanger. For one subsection, the properties of working fluid are assumed to be constant. The heat transfer equation for each subsection is given by

Q i ¼ U i  A i  Dt i

ð10Þ

where Dti is the log-mean temperature difference (LMTD) 3.1.2. Exergy analysis Exergy is the maximum theoretical work obtainable from an overall system consisting of a system and the environment as the system comes into equilibrium with the environment (passes to the dead state). The exergy flow rate of a fluid at kth state point is expressed as

Ek ¼ M k  ½ðhk  h0 Þ  T 0  ðsk  s0 Þ

ð7Þ

where the subscript 0 denotes the state of ambient conditions. The exergy rate balance equation for each component is written as

EF ¼ EP þ ED þ EL

ð8Þ

The detailed calculations of EF, EP, ED and EL for each component in the CCP system are listed in Table 2. The exergy efficiency is used to evaluate the thermodynamic performance of the system, which is given by

gexg ¼

W net þ Ecool ðEa  Ec Þ þ ðEd  Ee Þ

ð9Þ

where Ecool denotes the exergy rate of refrigeration process in the evaporator, which is equal to (E29  E28).



Dt i ¼

   T hf;iþ1  T cf ; i þ 1  T hf;i  T cf;i ln

ð11Þ

T hf;iþ1 T cf ;iþ1 T hf;i T cf;i

The total heat transfer coefficient for each subsection is calculated as

1 1 d 1 ¼ þ þ U i ahf;i k acf;i

ð12Þ

For single-phase flow, the convection heat transfer coefficient at the tube side is [34]

2

as;in ¼

kf 6 4 Din

3

12:7 



f 8

f 8 0:5

 Re  Pr 7 5  2  3  Pr  1 þ 1:07

ð13Þ

In Eq. (13), f, Re and Pr denote the Darcy friction factor, the Reynolds number and the Prandtl number, respectively, which are defined as

f ¼ ð0:790 ln Re  1:64Þ Re ¼

Gin Din

l

2

ð14Þ ð15Þ

307

J. Xia et al. / Energy Conversion and Management 128 (2016) 303–316 Table 2 Exergy analysis for each component in the CCP system. Component

EF

EP

ED

EL

Gas heater BC turbine Vapor generator 1 Precooler Compressor ORC turbine Condenser 1 Pump 1 Preheater Separator Vapor generator 2 Ejector Condenser 2 Valve Evaporator Pump 2

Ea  Eb E1  E2 E2  E3 / Wcomp E7  E8 / Wpump1 Ed  Ee E11 Eb  Ec E14 + E19 / / E18  E19 Wpump2

E1  E5 Wbt E6  E13 / E5  E4 Wot / E10  E9 E11  E10 E12 + E13 E14  E21 E15 / / E29  E28 E21  E20

Ea + E5  Eb  E1 E1  E2  Wbt E2 + E13  E3  E6 E3 + E22  E4  E23 Wcomp + E4  E5 E7  E8  Wot E8 + E24  E9  E25 Wpump1 + E9  E10 Ed + E10  Ee  E11 E11  E12  E13 Eb + E21  Ec  E14 E14 + E19  E15 E15 + E26  E16  E27 E17  E18 E18 + E28  E19  E29 Wpump2 + E20  E21

/ / / E23  E22 / / E25  E24 / / / / / E27  E26 / / /

Pr ¼

cp l k

ð16Þ

where Gin is the mass velocity of the flow inside the tube, being given by

Gin ¼

4M

ð17Þ

N  p  D2in

The convection heat transfer coefficient for single-phase at the shell side is calculated by Kern [35]

as;out

 0:55  0:14 k Des  u  q l ¼ 0:36    Pr1=3  Des lw l

ð18Þ

where Des is the equivalent diameter of the shell-side flow channel defined as

Des ¼

ð19Þ

The heat transfer process of the two-phase flow consists of the evaporation and the condensation processes. It is assumed that the organic fluid flows inside the tube for the vapor generators, condensers and evaporator, then the evaporation heat transfer coefficient is given by Gungor and Winterton [36]



0:8 Gð1  xÞd

ll

"



Pr0:4 l

Dip ¼

Q vs

ð23Þ

mt  1 4Aip 2

 x 0:75  ql 0:41 kl 1 þ 3000Bo0:86 þ 1:12 1x d qv

#

qm G  rf

ð24Þ

p

where Aip and Dip denote the cross sectional area and the diameter of the inlet pipe, respectively; Qvs is the volumetric flow; and mt is terminal velocity of the two phase flow, which is given by Zarrouka and Purnanto [38]

mt ¼ K

ql  qv qv

ð25Þ

In Eq. (25), K is a constant with a value of 0.069 ms1. In order to simplify the calculation, the vertical BOC separator is considered as a cylindrical vessel with a height of (LA + LB) and a diameter of Dsep, the calculation of the volume for the separator is given by

V sep ¼

p  D2sep  ðLA þ LB Þ 4

ð26Þ

ð20Þ

where LA is the length of the part which is above the inlet pipe, and LB is the length of the part which is below the inlet pipe. It is recommended by Ref. [38]: Dsep = 3Dip, LA = 7Dip and LB = 4.5Dip. After that, Eq. (26) could be written as

ð21Þ

V sep ¼

where Bo denotes the boiling number and expressed as

Bo ¼

Aip ¼



1:10P 2t  Dout Dout

atp;evp ¼ 0:023

3.2.1.2. Volume of the separator. Vertical BOC (bottoming outlet cyclone separator) separator is selected in this study, due to its high separator efficiency and simple structure. The calculations of the vessel dimensions are all given based on the two-phase inlet pipe. The pipe size is calculated as follows [38]

p  ð3Dip Þ2 4

 ð7Dip þ 4:5Dip Þ

ð27Þ

The condensation heat transfer coefficient is defined as [37]



atp;cnd ¼ 0:023 "



0:8 Gf ð1  xÞd

ll

Pr 0:4 l

kl 3:8x0:76 ð1  xÞ0:04 ð1  xÞ0:8 þ d p0:38 r

# ð22Þ

In Eq. (22), pr denotes the reduced pressure which is the ratio of state point pressure to critical pressure of the fluid. Eventually, the area of each heat exchanger can be obtained by adding up the areas for all subsections from both single-phase and two-phase regions.

3.2.2. Costs of equipments For the sake of estimating the equipment costs, the method of equipment module costing [39] is used in this paper. This method of costing relates all costs of the equipment to the purchased cost of equipment evaluated for some base conditions. The deviations from base conditions are handled by using multiplying factors depending on the specific equipment type, system pressure and materials of construction. Considering the effect of inflation, the CEPCI (Chemical Engineering Plant Cost Index) is employed to update the equipment costs to the year of 2014 [40], which is expressed as

308

C 2014

J. Xia et al. / Energy Conversion and Management 128 (2016) 303–316

  I2014 ¼ Cb  Ib

ð28Þ

where Cb refers to the cost at the base time; I2014 and Ib are the cost indexes assigned 576.1 and 397, respectively. The purchased cost of the equipment, at ambient pressures and using carbon steel construction, is calculated as [39]

log C 0eq ¼ K 1;eq þ K 2;eq  logðXÞ þ K 3;eq  ½logðXÞ

2

 576:1 0   C he  B1;he þ B2;he  F M;he  F P;he 397

ð30Þ

where B1,he and B2,he are the constants according to the type of the heat exchanger; FM,he is the material factor; FP,he is the pressure factor, which is expressed as 2

log F P;he ¼ C 1;he þ C 2;he  log Phe þ C 3;he  ðlog Phe Þ

ð31Þ

where C1,he, C2,he and C3,he are the constants according to the type and the pressure range of the heat exchanger. The turbines in this study are made from the material of carbon steel, with axial types, and the costs are calculated as

C turb ¼

576:1 0  C turb  F BM;turb 397

ð32Þ

where FBM,turb denotes the bare module factor based on the type and material of construction for the turbine. The compressor is axial type made from carbon material, and the cost is given by

C comp ¼

576:1 0  C comp  F BM;comp 397

ð33Þ

where FBM,comp is the bare module factor for the compressor. For the separator, the material of carbon steel and the vertical type are designed for the construction. The cost is expressed as

C sep ¼

 576:1 0   C sep  B1;sep þ B2;sep  F M;sep  F P;sep 397

ð34Þ

where B1,sep and B2,sep are the constants based on the type of the separator; FM,sep is the material factor; FP,sep is the pressure factor, which is calculated by

ðP sep þ1ÞDsep 2½8500:6ðP sep þ1Þ

þ 0:00315

0:0063

) ;1

ð35Þ

The pumps are designed to centrifugal type and made from stainless steel, and the costs of them are given by

C pump ¼

ð29Þ

where K1,eq, K2,eq and K2,eq are the constants for the equipment. And X is the size or capacity parameter for the equipment, representing the total heat transfer area of the heat exchanger (Ahe), the power generation by the turbine (Wturb), the volume of the separator (Vsep) and the power consumption by the pump (Wpump) or compressor (Wcomp). It is assumed that each heat exchanger of the system is made from the material of carbon steel, with a type of shell-and-tube. Then the cost is given by

C he ¼

( F P;sep ¼ max

  576:1 0  C pump  B1;pump þ B2;pump  F M;pump  F P;pump 397

ð36Þ

where B1,pump and B2,pump are the constants based on the type of the pump; FM,pump is the material factor; FP,pump is the pressure factor, which is given by

log F P;pump ¼ C 1;pump þ C 2;pump  log Ppump þ C 3;pump  2  log Ppump

ð37Þ

where C1,pump, C2,pump and C3,pump are the constants in terms of the type and the pressure range of the pump. The costs of the ejector, the valve and the working fluid are neglected in this paper, since their costs are much lower than those of other equipments [41,42]. The constants mentioned above are all listed in Table 3. 3.2.3. Exergoeconomic analysis In order to relate the present value of the expenditure to the equivalent annually levelized costs, the capital recovery factor (CRF) is used, which is calculated as [43] n

CRF ¼

ieff  ð1 þ ieff Þ n ð1 þ ieff Þ  1

ð38Þ

where ieff is the interest rate with a value of 0.05, and n is the lifetime of the CCP system with a value of 30 [44]. The equipment annually levelized costs is given by

Z i ¼ CRF  C i

ð39Þ

It is assumed that the annual working time of the CCP system is 8000 h [45]. The annual exergy transfer rates (Ey), annual power generation and consumption (Wy) are then obtained by this assumption. From the exergoeconomic point of view, each exergy flow is related to a cost, and the cost balance for kth component of the CCP system is given by Bejan et al. [43]

X X ðcout  Eout Þk þ cw;k  W k ¼ cq;k  Eq;k þ ðcin  Ein Þk þ Z k out

ð40Þ

in

where cout, cin, cw,k, and cq,k denote the average costs per unit of exergy. Note that the term (cw,k  Wk) would move with its positive sign to the right side of this equation if a component consumes power (such as the pump or compressor). The term (cq,k  Eq,k) would come out with its positive sign on the left side when there

Table 3 Constants for equipment costs calculation [39]. Constant

Value

Constant

Value

Constant

Value

K1,he K2,he K3,he K1,turb K2,turb K3,turb K1,comp K2,comp K3,comp K1,sep K2,sep K3,sep K1,pump

4.3247 0.3030 0.1634 2.7051 1.4398 0.1776 2.2897 1.3604 0.1027 3.4974 0.4485 0.1074 3.3892

K2,pump K3,pump B1,he B2,he B1,sep B2,sep B1,pump B2,pump C1,he (0.5 < Phe < 14) C2,he (0.5 < Phe < 14) C3,he (0.5 < Phe < 14) C1,he (Phe < 0.5) C2,he (Phe < 0.5)

0.0536 0.1538 1.63 1.66 2.25 1.82 1.89 1.35 0.03881 0.11272 0.08183 0 0

C3,he (Phe < 0.5) C1,pump (1 < Ppump < 10) C2,pump (1 < Ppump < 10) C3,pump (1 < Ppump < 10) C1,pump (Ppump < 1) C2,pump (Ppump < 1) C3,pump (Ppump < 1) FM,he FBM,turb FBM,comp FM,sep FM,pump

0 0.3935 0.3957 0.00226 0 0 0 1.0 3.5 2.7 1.0 2.2

J. Xia et al. / Energy Conversion and Management 128 (2016) 303–316 Table 4 Cost balance and auxiliary cost relations for each component [43]. Component

Cost balance

Auxiliary equation

Gas heater BC turbine Vapor generator 1 Precooler

c1  E1,y + cb  Eb,y = c5  E5,y + ca  Ea,y + Zgh c2  E2,y + cbt  Wbt,y = c1  E1,y + Zbt c3  E3,y + c6  E6,y = c2  E2,y + Zvp1

ca = cb c1 = c2 c2 = c3

c23  E23,y + c4  E4,y = c22  E22,y + c3  E3,y + Zpc c5  E5,y = c4  E4,y + ccomp  Wcomp,y + Zcomp c1  E8,y + cot  Wot,y = c7  E7,y + Zot c25  E25,y + c9  E9,y = c24  E24,y + c8  E8,y + Zcond1 c10  E10,y = c9  E9,y + cpump1  Wpump1,y + Zpump1 ce  Ee,y + c11  E11,y = cd  Ed,y + c10  E10,y + Zph c12  E12,y + c13  E13,y = c11  E11,y + Zsep c14  E14,y + cc  Ec,y = c21  E21,y + cb  Eb,y + Zvp2 c15  E15,y = c14  E14,y + c19  W19,y c27  E27,y + c16  E16,y = c26  E26,y + c15  E15,y + Zcond2 / c29  E29,y + c19  E19,y = c28  E28,y + c18  E18,y + Zevap c21  E21,y = c20  E20,y + cpump2  Wpump2,y + Zpump2

c3 = c4

Compressor ORC turbine Condenser 1 Pump 1 Preheater Separator Vapor generator2 Ejector Condenser 2 Valve Evaporator Pump 2

/ c7 = c8 c8 = c9 / cd = ce c12 = c13 cb = cc / c15 = c16 c17 = c18 c18 = c19 /

ð41Þ

cot ¼

c7  ðE7;y  E8;y Þ Z ot þ W ot;y W ot;y

ð42Þ

ð43Þ

where ccool denotes the average costs per unit of cold exergy produced by evaporator and is equal to c29. On the right side of the equations, the former part is defined as the fuel-exergy-related part, while the latter is considered as the equipment-cost-related part. In this paper, the average cost per unit of exergy product is used as an indicator for the system performance from the viewpoint of exergoeconomics, which is defined as



C diff;cond2 ¼ c27  E27;y  c26  E26;y ¼ c15  ðE15  E16 Þ þ Z cond2

ð47Þ

To ensure the accuracy of the calculation, validation for each cycle is considered based on the data from published literatures. The detailed comparisons between the present results and previous studies are listed in Tables 5–8. Note that the validation of ORC is divided into two parts, namely the separation part and the simple ORC part, respectively. And the working fluid of ORC is the same as that of ORC in the references only in this validation section. It can be seen from the tables that the simulation results in present study are in good agreement with the data from Refs. [49,50,11,51].

4. Results and discussion

  c1  E1;y  E2;y Z bt þ W bt;y W bt;y

cproduct ¼

ð46Þ

3.3. Validation

cbt ¼

c18  ðE18  E19 Þ Z evap þ E29 E29

C diff;cond1 ¼ c25  E25;y  c24  E24;y ¼ c8  ðE8  E9 Þ þ Z cond1

The simulation is performed under MATLAB software environment and REFPROP 9.0 [48] is employed to calculate the thermodynamic properties for the exhaust gases and the working fluids in this paper.

is a heat transfer from the component. Detailed cost balances and auxiliary cost relations for the components in CCP system are summarized in Table 4. The average costs per unit of exergy for the heat source of exhaust gases and jacket, the water entering to the condensers and evaporator are considered as zero [46]. Thus, the average costs per unit of exergy for each flow can be solved by resolving the linear equation system. In addition, taken the given conditions into account, the average costs per unit of exergy product for the turbines and evaporator could be written by

ccool ¼

In this study, a commercial cogeneration engine is selected as the case study. The main parameters of the engine are listed in Table 9. It is assumed that the composition of the exhaust gases are CO2, H2O, N2 and O2 with mass fraction of 9.1%, 7.4%, 74.2% and 9.3%, respectively [4]. This assumption can be used to evaluate the properties of the exhaust gases. Firstly, the thermodynamic performance of the proposed system is compared with those of the ones designed by Kim et al. [8] and Shu et al. [21], which focus on the waste heat recovery of engine exhaust and coolant. Then five key parameters such as the compressor pressure ratio, the compressor inlet temperature, the BC turbine inlet temperature, the ORC turbine inlet pressure and the ejector primary flow pressure, are selected to examine the thermodynamic and exergoeconomic performances of the CCP system. In the process of parametric analysis, when one parameter is varied, the others parameters are kept constant. The detailed conditions of the simulation are listed in Table 10. 4.1. Comparison of the proposed system and current systems Table 11 lists the comparison between the proposed system and the ones presented in Refs. [8,21]. It can be observed that total the net energy output (Wnet + Qcool) of proposed system is much higher than the maximum net energy output of the systems exhibited in the references under the same heat source conditions. Thus, the proposed system is more efficient to make better use of engine waste heat.

cwbt  W bt;y þ cwot  W ot;y þ ccool  Ecool;y þ C diff;pc þ C diff;cond1 þ C diff;cond2   W bt;y þ W ot;y þ Ecool;y

where Cdiff,pc, Cdiff,cond1 and Cdiff,cond2 denote the fictitious cost rates related to the utilization of dissipative components introduced by Ref. [47], which are expressed as

C diff;pc ¼ c23  E23;y  c22  E22;y ¼ c3  ðE3  E4 Þ þ Z pc

309

ð45Þ

 ð44Þ

4.2. Thermodynamic and exergoeconomic analyses Fig. 2 shows the effect of compressor pressure ratio on net power outputs, refrigeration capacity and exergy efficiency of the system.

310

J. Xia et al. / Energy Conversion and Management 128 (2016) 303–316

Table 5 Data comparison with Ref. [49] for the validation of Brayton cycle. T (°C)

Present

Ref.

P (MPa)

Present

Ref.

W (kJ kg1)

Present

Ref.

T1 T2 T4 T5

600 449.05 20.01 114.55

600 449.0 20 114.6

P1 P2 P4 P5

20 5.73 5.73 20

20 5.73 5.73 20

Wturb1 Wcomp

169.85 50.45

169.9 50.4

Table 6 Data comparison with Ref. [50] for the validation of separation part in ORC. State

11 12 13

T (°C)

H (kJ kg1)

P (MPa)

S (kJ kg1 K1)

M (kg s1)

Present

Ref.

Present

Ref.

Present

Ref.

Present

Ref.

Present

Ref.

163 163 163

163 163 163

0.6665 0.6665 0.6665

0.6665 0.6665 0.6665

990.14 2760.7 688.43

990 2761 688.7

2.664 6.724 1.972

2.664 6.725 1.973

1 0.1456 0.8544

1 0.1454 0.8546

Table 7 Data comparison with Ref. [11] for the validation of the conventional part ORC. P7 (MPa)

T7 (°C)

P9 (MPa)

T9 (°C)

M (kg s1)

Wturb2 (kW)

Wpump2 (kW)

gth (%)

R245fa

Present Ref.

1.4293 1.4293

107.75 107.75

0.1874 0.1874

31.29 31.29

0.5041 0.4988

10.615 10.615

0.621 0.615

8.40 8.38

R123

Present Ref.

1.5835 1.5835

132.85 132.85

192.6 192.6

46.85 46.85

0.5457 0.5458

10.652 10.652

0.652 0.644

8.88 8.6

Fluid

Table 8 Data comparison with Ref. [51] for the validation of ejector refrigeration cycle. H (kJ kg1)

P (MPa)

S (kJ kg1 K1)

M (kg s1)

State

T (°C) Present

Ref.

Present

Ref.

Present

Ref.

Present

Ref.

Present

Ref.

14 15 18 19

67.85 163 6.85 6.85

67.85 163 6.85 6.85

401 147.42 71.75 71.54

401 149 73 73

458.13 443.55 232.26 409.48

459.1 446.9 232.7 410.1

1.806 1.819 1.116 1.749

1.809 1.830 1.117 1.751

1 1.3175 0.3135 0.3135

1 1.3331 0.3331 0.3331

Table 9 Main parameters of engine [4].

Table 10 Simulation conditions of the CCP system.

Term

Value

Unit

Term

Value

Unit

Electrical power output Engine speed Exhaust gas temperature Exhaust mass flow Engine jacket temperatures Engine jacket flow

2928 1000 470 15,673 79/90 90

kW rpm °C kg/h °C m3/h

Environment temperature Environment pressure Compressor inlet pressure Compressor inlet temperature Compressor pressure ratio BC turbine inlet temperature ORC turbine inlet pressure Ejector primary flow inlet pressure Terminal temperature difference at gas heat outlet Terminal temperature difference at vapor generator 1 inlet Terminal temperature difference at vapor generator 2 inlet Pinch point temperature difference in vapor generator 1 Pinch point temperature difference in vapor generator 2 Condensation temperature of condenser 1 Condensation temperature of condenser 2 Evaporation temperature of evaporator Isentropic efficiency of BC turbine Isentropic efficiency of ORC turbine Isentropic efficiency of compressor Isentropic efficiency of pump 1 Isentropic efficiency of pump 2 Inlet temperature of cooling water Outlet temperature of cooling water in precooler Outlet temperature of cooling water in condenser 1 Outlet temperature of cooling water in condenser 2

15 0.1013 0.3 40 4 420 1.2 1.5 15 50 65 10 60 35 25 5 80 80 80 70 70 15 25 30 25

°C MPa MPa °C / °C MPa MPa °C °C °C °C °C °C °C °C % % % % % °C °C °C °C

It can be observed that net power output of Brayton cycle increases firstly and then decreases with the increasing compressor pressure ratio. The reason for this is that an increase in compressor pressure ratio leads to increases in enthalpy drop through the turbine and power consumption by pump. The former case can account for the initial rise of power output through the BC turbine. But as the effect of increasing power consumption through compressor gradually outweighs that of the increasing power produced by BC turbine, a decrease in net power output of BC occurs. The net power output of organic Rankine cycle decreases when the compressor pressure ratio increases. Since an increase in compressor pressure ratio results in a reduction in BC turbine outlet temperature, then the ORC turbine inlet temperature decreases. In addition, the mass flow rate of vapor generated by the vapor generator 1 also decreases due to the reducing BC turbine outlet temperature. Hence the net power output of ORC decreases. In ejector refrigeration cycle, the refrigeration capacity increases with the increasing pressure ratio of compressor. As

the compressor pressure ratio increases, the compressor outlet temperature rises under the condition of invariable compressor inlet temperature. Since the terminal temperature difference of

311

J. Xia et al. / Energy Conversion and Management 128 (2016) 303–316 Table 11 Comparison between the proposed system and current systems. Term

Ref. [8]

Present

Ref. [21]

Present

Engine power output (kW) Exhaust temperature (°C) Engine coolant temperature (°C) Exhaust mass flow rate (kg h1) Net power output (kW) Cooling capacity (kW) Total net energy output (kW)

26.7 689 100 118 5.63 0 5.63

26.7 689 100 118 6.38 0.96 7.34

235.8 519 83.3 990.79 36.77 0 36.77

235.8 519 83.3 990.79 33.79 7.52 41.31

WBC WORC Wnet

275 250

Qcool 30

exg

225

Output /kW

150 26

125 100

24

75 50

Exergy Efficiency /%

28

22

25 20

0 2.6

2.8

3.0

3.2

3.4

3.6

3.8

4.0

4.2

4.4

4.6

4.8

Pressure ratio Fig. 2. Effect of compressor pressure ratio on thermodynamic performance of the system.

Average cost per unit of exergy

/$·(MWh)-1

gas heater outlet is given as a constant, the exhaust gas temperature at vapor generator 2 inlet rises, leading to a higher mass flow rate of primary flows. Then the refrigeration capacity increases. Since the exergy output of refrigeration capacity is much smaller than that of net power production in terms of exergy analysis, the exergy efficiency of the system is more related to the net power output of the system. With the decreasing net power output of the system (Wnet), the exergy efficiency decreases. Fig. 3 illustrates the effect of compressor pressure ratio on average cost per unit of exergy product for the turbines, the evaporator and the overall system.

cproduct cwbt cwot ccool

2000 1600 1200 800 400

c1 c7 c18

125 100 75 50 25 0 2.8

3.0

3.2

3.4

3.6

3.8

4.0

4.2

4.4

4.6

4.8

Pressure ratio Fig. 3. Effect of compressor pressure ratio on exergoeconomic performance of the system.

It can be seen that the average cost per unit of exergy product for BC turbine rises with the increasing compressor pressure ratio. This can be explained by dividing that cost into two parts, namely the fuel-exergy-related part and the equipment-cost-related part from Eq. (41). For the fuel-exergy-related part, it can be seen from Fig. 3 that the average cost per unit of exergy fuel (c1) increases with rise of pressure ratio in the compressor, leading to an increase in the fuel-exergy-related part of cwbt. Referring to the equipmentcost-related part, the effect of the rise in BC turbine cost outweighs that of the BC turbine power output, resulting in an increase in equipment-cost-related part of cwbt. Combining the effects of these two parts, the average cost per unit of product for BC turbine increases. As the pressure ratio of compressor increases, the average cost per unit of exergy product for ORC turbine rises. Considering the fuel-exergy-related part, it can be observed from Fig. 3 that as the compressor pressure ratio increases, the average cost per unit of exergy fuel for ORC turbine rises (c7), thus an increase in the fuel-exergy-related part of cwot occurs. For the equipment-costrelated part, the impact of the increase in ORC turbine cost is greater than that of the ORC turbine power output, leading to a rise in equipment-cost-related part of cwot. As a result, the average cost per unit of product for ORC turbine increases. The average cost per unit of refrigeration capacity exergy output decreases with the increasing compressor pressure ratio. The reason for this is that an increase in compressor pressure ratio leads to a decrease in the average cost per unit of exergy fuel for the evaporator, as shown in Fig. 3, reducing the numerical value of fuel-exergy-related part of ccool. For the equipment-costrelated part, the impact of the decrease in evaporator cost is greater than that of the refrigeration capacity exergy output, resulting in a decrease in equipment-cost-related part of ccool. Therefore, a drop in average cost per unit of refrigeration capacity exergy output occurs. As for the fictitious cost rate of precooler (Cdiff,pc), both the exergy transfer rate and the equipment cost declines, due to the reduction of temperature at the BC turbine outlet caused by the increasing compressor pressure ratio. In addition, the average cost per unit of the exergy for the flow across the precooler (c3, which is equal to c1) increases, as is shown in Fig. 3. Integrating the variations of all the parameters related to Cdiff,pc in Eq. (45), the fictitious cost rate of precooler increases. Referring to the fictitious cost rate associated with condenser 1 (Cdiff,cond1), the reduced mass flow rate of working fluid in ORC leads to decreases in exergy transfer rate and equipment cost of condenser 1. Meanwhile, it can be observed that the average cost per unit of exergy for the flow across the condenser 1 (c8, which is equal to c7) rises with the increasing compressor pressure ratio. Thus the fictitious cost rate of condenser 1 increases. For the fictitious cost rate associated with condenser 2 (Cdiff,cond2), the higher mass flow rate of primary flows is generated by vapor generator 2 when compressor pressure ratio increases, causing the exergy transfer rate and equipment cost of condenser 2 to increase. Even though the average cost per unit of exergy for the flow across the condenser 2 (c15, which is equal to c18) decreases, the fictitious cost rate of condenser 2 still increases.

J. Xia et al. / Energy Conversion and Management 128 (2016) 303–316

The average cost per unit of exergy product of the overall system (cproduct) increases with the increasing compressor pressure ratio, after analyzing the variations for all parameters related to cproduct in Eq. (44). Fig. 4 shows the effect of compressor inlet temperature on net power outputs, refrigeration capacity and exergy efficiency of the system. It can be observed that the net power output of BC decreases with the increasing compressor inlet temperature. The reason is that as the compressor inlet temperature increases, the enthalpy increment in compressor and the mass flow rate of BC both increase, resulting in the increasing consumption of power through compressor. In the BC turbine, the power generation increases due to the raised mass flow rate of CO2. Combining these factors, the net power output of BC decreases. The net power output of ORC increases when the compressor inlet temperature rises. It is because an increase in mass flow rate of CO2 leads to a larger amount of vapor produced by vapor generator 1, causing the power output of ORC turbine to rise. Thus the net power output of ORC increases. In the ejector refrigeration cycle, the refrigeration capacity exergy output increases with the rising compressor inlet temperature. Since the rise of compressor inlet temperature enables the compressor outlet temperature to increase, the exhaust gases temperature at the gas heater outlet rises due to the constant terminal temperature difference at gas heater outlet. Therefore, more vapor is generated by vapor generator 2, contributing to an increase in the refrigeration capacity exergy output. Considering the variations of net power outputs and the refrigeration capacity exergy output of the system, the exergy efficiency decreases. Fig. 5 illustrates the effect of compressor inlet temperature on average cost per unit of exergy product for the turbines, the evaporator and the overall system. It can be seen that the average cost per unit of exergy product for BC turbine increases with the rising compressor inlet temperature. As the compressor inlet temperature rises, the fuel-exergyrelated part of cwbt increases due to the rised average cost per unit of exergy fuel (c1) for BC turbine. While for the equipment-costrelated part, the effect of the rise in BC turbine power output is greater than that of the BC turbine cost, leading to a decrease in equipment-cost-related part of cwbt. Integrating the effects of these

Qcool

WBC WORC Wnet

280 240

exg

30

Output /kW

26 140 24

120 100

Exergy Efficiency /%

28 160

22

80 34

36

38

40

42

44

20 46

T4/ Fig. 4. Effect of compressor inlet temperature on thermodynamic performance of the system.

Average cost per unit of exergy /$·(MWh)-1

312

cproduct cwbt cwot ccool

800 700 600 500 400

c1 c7 c18

80 60 40 20 0 34

36

38

40

42

44

46

T4/ Fig. 5. Effect of compressor inlet temperature on exergoeconomic performance of the system.

two parts, the average cost per unit of exergy product for BC turbine increases. As shown in Fig. 5, the average cost per unit of exergy fuel (c7) for ORC turbine increases when the compressor inlet temperature increases, resulting in an increase in fuel-exergy-related part of cwot. Referring to the equipment-cost-related part, the impact of the increase in ORC turbine power output outweighs that of the ORC turbine cost when the compressor inlet temperature rises, leading to a decrease in the equipment-cost-related part of cwot. As a result, the average cost per unit of exergy product for ORC turbine increases with the rising temperature of compressor inlet. With the rising compressor inlet temperature, the average cost per unit of product for the evaporator decreases. For the fuelexergy-related part, a decrease in the average cost per unit of exergy fuel (c18) leads to a decrease in the fuel-exergy-related part of ccool for the evaporator. Considering the equipment-cost-related part, since the impact of the rise in refrigeration capacity exergy output is greater than that of the increase in evaporator cost, the equipment-cost-related part of ccool decreases. Therefore, ccool decreases. Referring to the fictitious cost rate of precooler (Cdiff,pc), as the compressor inlet temperature increases, both the exergy transfer rate and the equipment cost for the precooler decrease. It can also be seen from the figure that the average cost per unit of exergy for the flow across the precooler (c3, which is equal to c1) increases slightly. Hence the fictitious cost rate of precooler decreases. For the fictitious cost rate of condenser 1 (Cdiff,cond1), a rise in mass flow rate of ORC leads to increases in both exergy transfer rate and the equipment cost of the condenser 1. Moreover, the average cost per unit of exergy for the flow across the condenser 1 (c8, which is equal to c7) increases. Thus the fictitious cost rate of the condenser 1 rises. Concerning the fictitious cost rate of condenser 2 (Cdiff,cond2), a larger mass flow rate of primary flow enables the exergy transfer rate and the equipment cost of the condenser 2 to increase. Whereas, the average cost per unit of exergy for the flow across the condenser 2 (c15, which is equal to c18) decreases. As a result, the fictitious cost rate of condenser 2 increases. After analyzing the variations of all parameters related to cproduct in Eq. (44), the average cost per unit of exergy product for the overall system increases with the increasing compressor inlet temperature. Fig. 6 shows the effect of BC turbine inlet temperature on net power outputs, refrigeration capacity and exergy efficiency of the system.

313

J. Xia et al. / Energy Conversion and Management 128 (2016) 303–316

26 exg

25

200

24

160

23

120

22

80

21

Exergy Efficiency /%

20

40 380

390

400

410

420

430

440

T1/ Fig. 6. Effect of BC turbine inlet temperature on thermodynamic performance of the system.

Average cost per unit of exergy /$·(MWh)-1

As BC turbine inlet temperature increases, the mass flow rate of CO2 decreases, leading to declines in both power generation through BC turbine and energy consumption by compressor. Since the impact of the decrease in power consumption by compressor is greater than that of the decrease in power generation through BC turbine, the net power output of BC increases. The net power output of ORC decreases when the BC turbine inlet temperature rises. Since the decrease in mass flow rate of CO2 leads to a decrease in mass flow rate of working fluid for ORC, the power output of ORC turbine decreases, causing the net power output of ORC to decrease. It can also be observed that the refrigeration capacity exergy output keeps constant with the increasing BC turbine inlet temperature. As BC turbine inlet temperature increases, the operation conditions of the compressor are not changed. Due to the constant terminal temperature difference of the gas heater outlet, the exhaust gases temperature stays unchanged at the vapor generator 2 inlet, resulting in the unchanged parameters of refrigeration cycle. Hence the refrigeration capacity exergy output is constant. Finally, the exergy efficiency of the overall system increases with the rise of BC turbine inlet temperature, due to the increased net power output of the system (Wnet). Fig. 7 illustrates the effect of BC turbine inlet temperature on average cost per unit of exergy product for the turbines, the evaporator and the overall system.

It can be seen that the average cost per unit of exergy product for BC turbine decreases with the increasing BC turbine inlet temperature. As the BC turbine inlet temperature rises, the fuelexergy-related part of cwbt decreases due to the reduced average cost per unit of exergy fuel (c1). Meanwhile, the rising BC turbine inlet temperature leads to a greater impact of the decrease in BC turbine cost than that of the BC turbine power output and hence decreases the equipment-cost-related part of cwbt. Combining the effects of these two parts, the average cost per unit of exergy product for BC turbine decreases. The average cost per unit of product for ORC turbine decreases with the increase of BC turbine inlet temperature. For the fuelexergy-related part, as the BC turbine inlet temperature rises, the average cost per unit of exergy fuel (c7) decreases, resulting in a decrease of the fuel-exergy-related part for cwot. Considering the equipment-cost-related part, the degree of reduction for the ORC turbine cost outweighs that for ORC turbine power output when the BC turbine inlet temperature rises, thus the equipment-costrelated part of cwot decreases. As a result, the average cost per unit of exergy product for ORC turbine decreases. Meanwhile, the average cost per unit of refrigeration capacity exergy output keeps constant when the BC turbine inlet temperature increases, for the reason that the parameter conditions of ejector are not changed. Referring to the fictitious cost rate of precooler (Cdiff,pc), as the BC turbine inlet temperature increases, the mass flow rate of CO2 decreases, leading to decreases in both exergy transfer rate and equipment cost of the precooler. In addition, the average cost per unit of exergy for the flow in the precooler (c3, which is equal to c1) drops, as is shown in Fig. 7. Thus Cdiff,pc decreases. For the fictitious cost rate of condenser 1 (Cdiff,cond1), on one hand, the exergy transfer rate and the equipment cost decrease due to the drop in mass flow rate of working fluid. On the other hand, the average cost per unit of exergy for the flow in the condenser 1 (c8, which is equal to c7) decreases, as shown in Fig. 7. Hence the increase of Cdiff,cond1 occurs. Considering the fictitious cost rate of condenser 2 (Cdiff,cond2), Cdiff,cond2 is a constant because of the unchanged parameters in condenser 2. Integrating the effects of all the variables related to cproduct in Eq. (44), the average cost per unit of exergy product for the overall system decreases when BC turbine inlet temperature increases. Fig. 8 shows the effect of ORC turbine inlet pressure on net power outputs, refrigeration capacity and exergy efficiency of the system.

WBC WORC Wnet

300

600 560

cproduct cwbt cwot ccool

520

80

c1 c7 c18

60 40

250

Qcool

30

exg

25

200

Output /kW

Output /kW

240

Qcool

20 150 15

100 50

Exergy Efficiency /%

WBC WORC Wnet

280

10

20 0 5

0 380

390

400

410

420

430

440

450

T1/ Fig. 7. Effect of BC turbine inlet temperature on exergoeconomic performance of the system.

0.4

0.6

0.8

1.0

1.2

1.4

P6/MPa Fig. 8. Effect of ORC turbine inlet pressure on thermodynamic performance of the system.

J. Xia et al. / Energy Conversion and Management 128 (2016) 303–316

increases slightly. Consequently, Cdiff,pc increases. For the fictitious cost rate of condenser 1 (Cdiff,cond1), Cdiff,cond1 decreases mainly because of the decrease in the average cost per unit of exergy for the flow across the condenser 1 (c8, which is equal to c7). Since the parameter conditions of the condenser 2 is unchanged, the fictitious cost rate of condenser 2 (Cdiff,cond2) keeps constant. Due to the combined effects of all the variables related to cproduct in Eq. (44), the average cost per unit of exergy product for the overall system decreases with the increasing ORC turbine inlet pressure. Fig. 10 shows the effect of ejector primary flow pressure on net power outputs, refrigeration capacity and exergy efficiency of the system. It can be observed the net power outputs of BC and ORC keep as constants, for the reason that the parameters are irrelevant to the primary flow pressure of the ejector. Since a higher ejector primary flow pressure leads to a lower mass flow rate of the primary flow across the ejector, casing a drop in the mass flow rate of the secondary flow in the evaporator. The refrigeration capacity decreases. The exergy transfer rate in vapor generator 2 decreases with the increasing ejector primary flow pressure. Considering the variations of net power output, the refrigeration capacity exergy output and the exergy transfer rate of the system, the exergy efficiency slightly increases. Fig. 11 illustrates the effect of ejector primary flow pressure on the average cost per unit of exergy product for the turbines, the evaporator and the overall system. As is shown in the figure, the average costs per unit of exergy product for the BC turbine and the ORC turbine are both constants, due to the unchanged operation conditions of them. Meanwhile, the average cost per unit of exergy product for the evaporator increases with the rising ejector primary flow pressure. As the ejector primary flow pressure increases, the average cost per unit of exergy fuel (c18) for the evaporator rises, resulting in an increase in the fuel-exergy-related part of ccool. While for the equipment-cost-related part, the impact of the increase in evaporator cost outweighs that of the evaporator refrigeration capacity exergy output, leading to an increase in the equipment-costrelated part of ccool. Therefore, ccool increases. For the fictitious cost rates of the precooler and condenser 1, both Cdiff,pc and Cdiff,cond1 keep as constants since the operation conditions of them are not changed. While Cdiff,cond2 decreases by the reason of the decreased mass flow rate of working fluid across the condenser 2.

WBC WORC Wnet

280

600

260

550

Qcool 30

exg

240

500

cproduct cwbt cwot ccool

450 140 120 100

c1 c7 c18

80 60

28

220

Output /kW

Average cost per unit of exergy /$·(MWh)-1

It can be obtained from the figure that both the net power output of BC and the refrigeration capacity of the ejector refrigeration cycle remain as constants. Since the parameters of the BC turbine, compressor and the evaporator are unrelated to the inlet pressure of ORC turbine. An increase in ORC turbine inlet pressure leads to a decrease in the mass flow rate of working fluid and an increase in enthalpy through the turbine. Since the impact of the increase in enthalpy drop is greater than that of the decrease in the mass flow rate of working fluid across the ORC turbine, the power output of ORC turbine rises. Consequently, the net power output of ORC increases. Integrating the effects of the net power output and the refrigeration capacity exergy, the exergy efficiency increases with the increasing ORC turbine inlet pressure. Fig. 9 illustrates the effect of ORC turbine inlet pressure on the average cost per unit of exergy product for the turbines, the evaporator and the overall system. It can be seen that the average cost per unit of exergy product for BC turbine increases slightly with the increasing ORC turbine inlet pressure. Since a rise in the average cost per unit of exergy fuel for the BC turbine (c1) leads to an increase in the fuelexergy-related part of cwbt, and the equipment-cost-related part of cwbt is not varied due to the constant parameters of BC turbine. The average cost per unit of exergy product for BC turbine increases. The average cost per unit of exergy product ORC turbine decreases with the rising ORC turbine inlet pressure. For the fuelexergy-related part, as the ORC turbine inlet pressure rises, the average cost per unit of exergy fuel for the ORC turbine (c7) decreases, resulting in a decrease in the fuel-exergy-related part of cwot. Referring to the equipment-cost-related part, since the impact of the rise in ORC turbine power output is greater than that of the increase in ORC turbine cost, the equipment-cost-related part of cwot also decreases. Hence the average cost per unit of exergy product for ORC turbine decreases. With the increasing ORC turbine inlet pressure, the average cost per unit of exergy product for the evaporator remains unchanged, due to the constant average cost per unit of exergy fuel (c18) and the unchanged parameters for the evaporator. Referring to the fictitious cost rate of precooler (Cdiff,pc), as the ORC turbine inlet pressure increases, the temperature of saturated liquid at state point 13 rises, leading to an increase in the temperature at the precooler inlet. Then the exergy transfer and the equipment cost both increase. In addition, the average cost per unit of exergy for the flow across the precooler (c3, which is equal to c1)

26

160 140

24

120 100

40

22

80

20

Exergy Efficiency /%

314

60

0 0.4

0.6

0.8

1.0

1.2

1.4

1.6

P6/MPa Fig. 9. Effect of ORC turbine inlet pressure on exergoeconomic performance of the system.

1.4

1.6

1.8

2.0

2.2

2.4

20 2.6

P14/MPa Fig. 10. Effect of ejector primary flow pressure on thermodynamic performance of the system.

315

Average cost per unit of exergy /$·(MWh)-1

J. Xia et al. / Energy Conversion and Management 128 (2016) 303–316

cproduct cwbt cwot ccool

1000 800 600

c1 c7 c18

Table 13 Control parameters of GA. Tuning parameters

Value

Population size Crossover probability Mutation probability Stop generation

20 0.8 0.01 200

80 60

Table 14 Single-objective optimization results.

40 20 0 1.4

1.6

1.8

2.0

2.2

2.4

2.6

P14/MPa Fig. 11. Effect of ejector primary flow pressure on exergoeconomic performance of the system.

Combining the effects of all parameters related to cproduct in Eq. (44), the average cost per unit of exergy product for the overall system increases with the increasing ejector primary flow pressure. In order to obtain the optimal performance of the CCP system utilized engine waste heat, a single-objective optimization is conducted by means of genetic algorithm, which is a stochastic global search method that simulates natural biological evolution. The detailed computing method of genetic algorithm can refer to Ref. [3]. The average cost per unit of exergy product for the overall system is selected the as the objective function, since it combines exergy analysis and economic principles to provide information which is critical to the design and operation of a cost-effective system. Key parameters such as compressor pressure ratio, compressor inlet pressure, BC turbine inlet temperature, ORC turbine inlet pressure and ejector primary flow pressure are chosen as the decision variables. The detailed ranges of key parameters and the control parameters of GA are listed in Tables 12 and 13, respectively. Table 14 lists the results of parameter optimization for the CCP system with the objective function of the average cost per unit of exergy product for the overall system. It can be observed that the minimum cproduct is 63.53 $/MW h1, and the exergy efficiency, the net power output and the Refrigeration capacity of the system are 27.63%, 282.49 kW and 20.01 kW, respectively. The results indicate that the exergy efficiency obtained from the exergoeconimic optimization is also desirable, and the refrigeration capacity is optimized to be lower due to the higher cost per unit of exergy product for ejector refrigeration cycle. In addition, for the compressor pressure ratio, the optimal one is located at the lower boundary. Referring to the compressor inlet temperature and the ejector primary flow pressure, the optimum ones are settled near the lower boundaries. Concerning the optimal BC turbine inlet temperature and the ORC turbine inlet pressure, both of them

Term

Value

Unit

Compressor pressure ratio Compressor inlet temperature BC turbine inlet temperature ORC turbine inlet pressure Ejector primary flow pressure Net power output Refrigeration capacity Exergy efficiency Average cost rate per unit of exergy product

3 36.242 435.404 1.363 1.556 282.49 20.01 27.63 63.53

/ °C °C MPa MPa kW kW % $ (MW h)1

are located near the upper boundaries. These results imply that all the optimal values of the key parameters are in accordance with the parametric analysis.

5. Conclusion In this study, a CCP system is designed for the waste heat recovery of an ICE. The effects of five key parameters (i.e. compressor pressure ratio, compressor inlet pressure, BC turbine inlet temperature, ORC turbine inlet pressure and ejector primary flow pressure) on thermodynamic and exergoeconomic performances are examined. Then a single-objective optimization by means of GA is conducted with an objective function of the average cost per unit of exergy product. The main conclusions drawn from investigation are summarized as follows: (1) On the basis of parametric analysis, the increases of the BC turbine inlet temperature, the ORC turbine inlet pressure and the ejector primary flow pressure have positive effects on the exergy efficiency, while the rises in compressor pressure ratio and the compressor inlet temperature are undesirable to the thermodynamic performance of the system. (2) Referring to the exergoeconimic analysis, lower average cost per unit of exergy product for the overall system can be achieved by increasing the BC turbine inlet temperature, the ORC turbine inlet pressure and the ejector primary flow pressure. However, the rises in compressor pressure and compressor inlet temperature lead to increases in average cost per unit of exergy product for the overall system. (3) By the single-objective optimization, the lowest average cost per unit of exergy product for the overall system is obtained, and the exergy efficiency is also desirable. In addition, all the optimal values of the key parameters are in accordance with the parametric analysis.

Table 12 Ranges of the decision variables. Decision variables

Range

Compressor pressure ratio Compressor inlet temperature (°C) BC turbine inlet temperature (°C) ORC turbine inlet pressure (MPa) Ejector primary flow pressure

3–4.6 35–45 380–440 0.5–1.4 1.5–2.5

Acknowledgments The authors gratefully acknowledge the financial support by the National Natural Science Foundation of China (Grant No. 51476121) and the Fundamental Research Funds for the Central Universities (Grant No. 2013jdgz14).

316

J. Xia et al. / Energy Conversion and Management 128 (2016) 303–316

References [1] Domingues A, Santos H, Costa M. Analysis of vehicle exhaust waste heat recovery potential using a Rankine cycle. Energy 2013;49:71–85. [2] Sprouse C, Depcik C. Review of organic Rankine cycles for internal combustion engine exhaust waste heat recovery. Appl Therm Eng 2013;51:711–22. [3] Dai YP, Wang JF, Gao L. Parametric optimization and comparative study of organic Rankine cycle (ORC) for low grade waste heat recovery. Energy Convers Manage 2009;50:576–82. [4] Vaja I, Gambarotta A. Internal combustion engine (ICE) bottoming with organic Rankine cycles (ORCs). Energy 2010;35:1084–93. [5] Tahani M, Javan S, Biglari M. A comprehensive study on waste heat recovery from internal combustion engines using organic rankine cycle. Therm Sci 2013;17:611–24. [6] He MG, Zhang XX, Zeng K, Gao K. A combined thermodynamic cycle used for waste heat recovery of internal combustion engine. Energy 2011;36:6821–9. [7] Wang EH, Zhang HG, Zhao Y, Fan BY, Wu YT, Mu QH. Performance analysis of a novel system combining a dual loop organic Rankine cycle (ORC) with a gasoline engine. Energy 2012;43:385–95. [8] Kim YM, Shin DG, Kim CG, Cho GB. Single-loop organic Rankine cycles for engine waste heat recovery using both low- and high-temperature heat sources. Energy 2016;96:482–94. [9] Liu BT, Chien KH, Wang CC. Effect of working fluids on organic Rankine cycle for waste heat recovery. Energy 2004;29:1207–17. [10] Tian H, Shu GQ, Wei HQ, Liang XY, Liu LN. Fluids and parameters optimization for the organic Rankine cycles (ORCs) used in exhaust heat recovery of internal combustion engine (ICE). Energy 2012;47:125–36. [11] Wang EH, Zhang HG, Fan BY, Ouyang MG, Zhao Y, Mu QH. Study of working fluid selection of organic Rankine cycle (ORC) for engine waste heat recovery. Energy 2011;36:3406–18. [12] Li YR, Du MT, Wu CM, Wu SY, Liu C. Potential of organic Rankine cycle using zeotropic mixtures as working fluids for waste heat recovery. Energy 2014;77:509–19. [13] Shu GQ, Gao YY, Tian H, Wei HQ, Liang XY. Study of mixtures based on hydrocarbons used in ORC (organic Rankine cycle) for engine waste heat recovery. Energy 2014;74:428–38. [14] Pierobon L, Kandepu R, Haglind F. ASME. Waste heat recovery for offshore applications. New York: American Society of Mechanical Engineers; 2013. [15] Yu GP, Shu GQ, Tian H, Wei HQ, Liu LN. Simulation and thermodynamic analysis of a bottoming organic Rankine cycle (ORC) of diesel engine (DE). Energy 2013;51:281–90. [16] Miller EW, Hendricks TJ, Wang H, Peterson RB. Integrated dual-cycle energy recovery using thermoelectric conversion and an organic Rankine bottoming cycle. Proc Inst Mech Eng Part A-J Power Energy 2011;225:33–43. [17] Zhang CY, Shu GQ, Tian H, Wei HQ, Liang XY. Comparative study of alternative ORC-based combined power systems to exploit high temperature waste heat. Energy Convers Manage 2015;89:541–54. [18] Zhang HG, Wang EH, Fan BY. A performance analysis of a novel system of a dual loop bottoming organic Rankine cycle (ORC) with a light-duty diesel engine. Appl Energy 2013;102:1504–13. [19] Yang FB, Dong XR, Zhang HG, Wang Z, Yang K, Zhang J, et al. Performance analysis of waste heat recovery with a dual loop organic Rankine cycle (ORC) system for diesel engine under various operating conditions. Energy Convers Manage 2014;80:243–55. [20] Choi BC, Kim YM. Thermodynamic analysis of a dual loop heat recovery system with trilateral cycle applied to exhaust gases of internal combustion engine for propulsion of the 6800 TEU container ship. Energy 2013;58:404–16. [21] Shu GQ, Liu LN, Tian H, Wei HQ, Yu GP. Parametric and working fluid analysis of a dual-loop organic Rankine cycle (DORC) used in engine waste heat recovery. Appl Energy 2014;113:1188–98. [22] Song J, Gu CW. Performance analysis of a dual-loop organic Rankine cycle (ORC) system with wet steam expansion for engine waste heat recovery. Appl Energy 2015;156:280–9. [23] Zhou YD, Wu YD, Li F, Yu LJ. Performance analysis of zeotropic mixtures for the dual-loop system combined with internal combustion engine. Energy Convers Manage 2016;118:406–14. [24] Liang YC, Shu GQ, Tian H, Wei HQ, Liang XY, Liu LN, et al. Theoretical analysis of a novel electricity-cooling cogeneration system (ECCS) based on cascade use of waste heat of marine engine. Energy Convers Manage 2014;85:888–94. [25] Wang S, Bai K, Xie Y, Di J, Cheng S. Analysis of combined power and refrigeration generation using the carbon dioxide thermodynamic cycle to

[26]

[27]

[28]

[29] [30]

[31] [32]

[33]

[34] [35] [36]

[37] [38] [39] [40]

[41]

[42]

[43] [44] [45]

[46]

[47]

[48]

[49] [50] [51]

recover the waste heat of an internal combustion engine. Math Probl Eng 2014;2014:1–12. Moisseytsev A, Sienicki JJ. Investigation of alternative layouts for the supercritical carbon dioxide Brayton cycle for a sodium-cooled fast reactor. Nucl Eng Des 2009;239:1362–71. Song BY, Zhuge WL, Zhao RC, Zheng XQ, Zhang YJ, Yin Y, et al. An investigation on the performance of a Brayton cycle waste heat recovery system for turbocharged diesel engines. J Mech Sci Technol 2013;27:1721–9. Akbari AD, Mahmoudi SMS. Thermoeconomic analysis & optimization of the combined supercritical CO2 (carbon dioxide) recompression Brayton/organic Rankine cycle. Energy 2014;78:501–12. Besagni G, Mereu R, Inzoli F. Ejector refrigeration: a comprehensive review. Renew Sust Energy Rev 2016;53:373–407. Shu GQ, Li XN, Tian H, Liang XY, Wei HQ, Wang X. Alkanes as working fluids for high-temperature exhaust heat recovery of diesel engine using organic Rankine cycle. Appl Energy 2014;119:204–17. Lai NA, Wendland M, Fischer J. Working fluids for high-temperature organic Rankine cycles. Energy 2011;36:199–211. Butrymowicz D, Smierciew K, Karwacki J, Gagan J. Experimental investigations of low-temperature driven ejection refrigeration cycle operating with isobutane. Int J Refrig-Rev Int Froid 2014;39:196–209. Zhang J, Zhang HG, Yang K, Yang FB, Wang Z, Zhao GY, et al. Performance analysis of regenerative organic Rankine cycle (RORC) using the pure working fluid and the zeotropic mixture over the whole operating range of a diesel engine. Energy Convers Manage 2014;84:282–94. Incropera FP, DeWitt DP. Fundamentals of heat and mass transfer. New York: Wiley; 2002. Kern DQ. Process heat transfer. New York: McGraw-Hill; 1950. Gungor KE, Winterton RHS. Simplified general correlation for saturated flow boiling and comparisons of correlations with data. Chem Eng Res Des 1987;65:148–56. Shah MM. A general correlation for heat transfer during film condensation inside pipes. Int J Heat Mass Transf 1979;22:547–56. Zarrouka SJ, Purnanto MH. Geothermal steam-water separators: design overview. Geothermics 2015;53:236–54. Turton R. Analysis, synthesis, and design of chemical processes. Upper Saddle River (NJ): Prentice Hall; 2009. Lozowski D. Start online account with the exclusive online plant cost index and get full archive access to chemical engineering. Chemical Engineering. http://www.chemengonline.com/pci-home. Farshi LG, Mahmoudi SMS, Rosen MA. Exergoeconomic comparison of double effect and combined ejector-double effect absorption refrigeration systems. Appl Energy 2013;103:700–11. Yang MH, Yeh RH. Economic performances optimization of the transcritical Rankine cycle systems in geothermal application. Energy Convers Manage 2015;95:20–31. Bejan A, Tsatsaronis G, Moran MJ. Thermal design and optimization. New York: Wiley; 1996. Tempesti D, Fiaschi D. Thermo-economic assessment of a micro CHP system fuelled by geothermal and solar energy. Energy 2013;58:45–51. Velez F, Segovia JJ, Martin MC, Antonlin G, Chejne F, Quijano A. A technical, economical and market review of organic Rankine cycles for the conversion of low-grade heat for power generation. Renew Sust Energy Rev 2012;16:4175–89. Sadeghi M, Mahmoudi SMS, Saray RK. Exergoeconomic analysis and multiobjective optimization of an ejector refrigeration cycle powered by an internal combustion (HCCI) engine. Energy Convers Manage 2015;96:403–17. Lazzaretto A, Tsatsaronis G. SPECO: a systematic and general methodology for calculating efficiencies and costs in thermal systems. Energy 2006;31:1257–89. Lemmon EW, Huber ML, McLinden MO. NIST standard reference database 23, reference fluid thermodynamic and transport properties (REFPROP). version 90. National Institute of Standards and Technology; 2010. Kim YM, Kim CG, Favrat D. Transcritical or supercritical CO2 cycles using both low- and high-temperature heat sources. Energy 2012;43:402–15. Yari M. Exergetic analysis of various types of geothermal power plants. Renew Energy 2010;35:112–21. Zheng B, Weng YW. A combined power and ejector refrigeration cycle for low temperature heat sources. Sol Energy 2010;84:784–91.